You are on page 1of 23

Pure Appl. Geophys.

168 (2011), 2097–2119


Ó 2011 Springer Basel AG (outside the USA)
DOI 10.1007/s00024-011-0291-5 Pure and Applied Geophysics

Nearshore Tsunami Inundation Model Validation: Toward Sediment Transport Applications


ALEX APOTSOS,1 MARK BUCKLEY,2 GUY GELFENBAUM,1 BRUCE JAFFE,2 and DEEPAK VATVANI3

Abstract—Model predictions from a numerical model, tsunamis have the potential to produce similar
Delft3D, based on the nonlinear shallow water equations are
compared with analytical results and laboratory observations from
disasters in the future. In order to develop appropriate
seven tsunami-like benchmark experiments, and with field obser- coastal management strategies that properly plan for
vations from the 26 December 2004 Indian Ocean tsunami. The and mitigate losses owing to tsunamis, researchers
model accurately predicts the magnitude and timing of the mea-
have generally focused on two approaches. The first
sured water levels and flow velocities, as well as the magnitude of
the maximum inundation distance and run-up, for both breaking involves coupling early detection systems with
and non-breaking waves. The shock-capturing numerical scheme numerical models that predict, in real-time, areas
employed describes well the total decrease in wave height due to most likely to be impacted by a tsunami (e.g., WEI
breaking, but does not reproduce the observed shoaling near the
break point. The maximum water levels observed onshore near et al., 2008; TANG et al., 2009). The second involves
Kuala Meurisi, Sumatra, following the 26 December 2004 tsunami improving local tsunami hazard assessments through
are well predicted given the uncertainty in the model setup. The a better understanding of the recurrence interval and
good agreement between the model predictions and the analytical
results and observations demonstrates that the numerical solution
magnitude of historic and pre-historic tsunamis (e.g.,
and wetting and drying methods employed are appropriate for GONZÁLEZ et al., 2009). Both of these approaches
modeling tsunami inundation for breaking and non-breaking long require a detailed understanding of the physical pro-
waves. Extension of the model to include sediment transport may
cesses that occur during a tsunami, particularly the
be appropriate for long, non-breaking tsunami waves. Using
available sediment transport formulations, the sediment deposit dynamics of nearshore propagation and inundation.
thickness at Kuala Meurisi is predicted generally within a factor The second approach also requires a detailed under-
of 2. standing of how tsunamis transport and deposit
Key words: Tsunami, inundation, run-up, benchmark valida- sediment. Here, we validate a hydrodynamic model
tion, numerical modeling. and demonstrate that extending it to include sediment
transport produces reasonable results that compare
well with field observations.
Unfortunately, the infrequent and unpredictable
1. Introduction nature of tsunamis makes it difficult to collect detailed
measurements of the physical processes that occur.
The 26 December 2004 tsunami in the Indian Therefore, to better understand tsunami processes,
Ocean killed over 237,000 people and left more than analytical, laboratory, and numerical experiments have
a million homeless (U.S. Agency for International been conducted. While much of the early work focused
Development (USAID), 2005). With a large segment on analytical solutions to simplified problems
of the world’s population living along coasts, major (e.g., CARRIER and GREENSPAN, 1958; TADEPALLI and
SYNOLAKIS, 1994, 1996; CARRIER et al., 2003; KÁNOGLU,
2004) and laboratory experiments employing solitary
waves (e.g., SYNOLAKIS, 1987; BRIGGS et al., 1995,
1
Pacific Coastal and Marine Science Center, USGS, 345 1996), recent increases in processing power have
Middlefield Rd, MS 999, Menlo Park, CA, USA. E-mail: greatly improved the utility of numerical models.
aapotsos@usgs.gov These models can now be used to predict the propa-
2
Pacific Coastal and Marine Science Center, USGS, 400
Natural Bridges Drive, Santa Cruz, CA, USA. gation and inundation paths of tsunamis in almost real-
3
Deltares, Delft, The Netherlands. time (e.g., TITOV et al., 2005; WEI et al., 2008; TANG
2098 A. Apotsos et al. Pure Appl. Geophys.

et al., 2009). The use of numerical models in this ocean. However, while analyses of video recordings
operational sense has allowed for the rapid determi- have provided a few estimates of onshore tsunami
nation of which communities are at risk once a tsunami flow velocities (FRITZ et al., 2006), detailed obser-
has been detected, and has helped coastal planners and vations in the nearshore and on land, where sediment
emergency responders act accordingly. transport is important, are limited. Model evaluation
Numerical models can also play a role in for these regions is therefore still primarily done
improving tsunami hazard assessments, which are using analytical and laboratory experiments.
often constrained by our limited knowledge of the Over the past two decades attempts have been
historic and pre-historic tsunami record. In order to made to standardize the process of model evaluation
estimate the probability that a major tsunami will to ensure the accuracy of model predictions. To this
strike a specific location, information is needed end, a set of benchmarks was developed during the
concerning tsunamis that occurred hundreds or International Workshops on Long-Wave Run-up
thousands of years ago. Typically the only evidence (e.g., YEH et al., 1996; LIU et al., 2008), and a pro-
that remains from these paleo-tsunamis is preserved cedure for model verification was outlined
in the geologic record as a sedimentary layer (e.g., (SYNOLAKIS et al., 2007, 2008). Following the estab-
ATWATER, 1987). Dating the soil layers surrounding lishment of these initial benchmarks, several other
these deposits can help identify approximately when useful laboratory experiments have been conducted
the tsunami occurred, and thus increase our under- and the data made available (see BALDOCK et al.,
standing of the recurrence interval of major tsunamis. 2008; SWILGER, 2009). However, while detailed
To improve hazard assessments, knowledge of the comparisons with laboratory observations remain a
hydrodynamic characteristics of the tsunami is also widely used method to test and validate numerical
necessary. Therefore, significant efforts have been tsunami models, especially in the nearshore, the use
made to use the information (e.g., sediment size, of laboratory observations is not without problems
deposit thickness, grading) preserved in tsunami (see MADSEN et al., 2008 and Sect. 4.1 in this paper),
sediment deposits to estimate the wave height, flow and offers only qualified model validation.
depth, and velocity of paleo-tsunamis (e.g., JAFFE and Here the results from a hydrodynamic numerical
GELFENBAUM, 2007; SMITH et al., 2007; NANAYAMA model, Delft3D, are compared with seven tsunami-
et al., 2007). These efforts, as well as efforts to like benchmarks, and with maximum water level
understand the scouring that occurs around obstacles observations measured near Kuala Meurisi, Sumatra,
during tsunamis, would benefit from a better under- following the 26 December 2004 Indian Ocean tsu-
standing of how sediment is transported and nami (JAFFE et al., 2006). The benchmarks include:
deposited by a tsunami. Numerical models that (1) the analytical solution of run-up on a plane beach
accurately predict the hydrodynamic processes of a (CARRIER et al., 2003), and laboratory experiments
tsunami could be coupled with sediment suspension using (2) a piece-wise linear bathymetry (BALDOCK
and transport formulations and used to determine et al., 2008), (3) an alongshore non-uniform beach
which processes dominate tsunami sediment transport (SWILGER, 2009), (4) a circular island amid a constant
and what information concerning the source mecha- depth basin (BRIGGS et al., 1995), (5) an alongshore
nism and hydrodynamic characteristics of the tsunami non-uniform beach with a circular island (http://
is retained in tsunami sediment deposits. isec.nacse.org/workshop/isec_workshop_2009), (6) a
Before a hydrodynamic model can be coupled complex three-dimensional beach (Monai Valley,
with sediment transport formulations, it must be Japan as described in SYNOLAKIS et al. (2008), and (7)
shown to depict accurately the hydrodynamic pro- a piece-wise linear bathymetry ending in a vertical
cesses of a tsunami. Recent deployment of DART wall (BRIGGS et al., 1996).
(Deep-Ocean Assessing and Reporting of Tsunamis) Our primary goal in this paper is to demonstrate,
underwater buoys (e.g., GONZALEZ et al., 2005) and to the extent possible, that Delft3D is appropriate for
the use of satellite imagery (e.g., ARCAS and TITOV, modeling the hydrodynamic processes that occur
2006) have improved model validation in the deep during the nearshore propagation and inundation of a
Vol. 168, (2011) Tsunami Model Validation: Toward Sediment Transport 2099

tsunami, and to examine the appropriateness of incompressibility of the flow. The extension of the
including sediment transport in the model. Verifica- model to three dimensions is important for many
tion that Delft3D accurately predicts the applications where resolution of the bottom boundary
hydrodynamic processes that occur during inundation layer or vertical stratification within the flow is
is advantageous because the model includes a important (i.e., see Sect. 4.2).
sophisticated sediment transport model that has been Numerous methods have been developed to solve
extensively validated in a wide variety of coastal the NLSWEs. The numeric solver used here is based
settings (LESSER et al., 2004; VAN RIJN et al., 2007). on the conservation of mass, momentum (during flow
This transport model allows sediment particles of expansions), and energy head (during flow contrac-
varying sizes to be transported in bed or suspended tions) (STELLING and DUINMEIJER, 2003). This method,
load, the vertical and horizontal structures of any referred to as the Flood Solver within the Delft3D
onshore sediment deposits to be determined, and the framework, was developed specifically to handle
morphological change associated with sediment rapidly varying flows, accurately predict the rapid
transport to be predicted. These features position the wetting and drying of grid cells, and be applicable to
model well to be a powerful tool for improving our a wide range of Froude numbers (including subcriti-
understanding of how tsunamis transport and deposit cal and supercritical flows). This numerical method
sediment. has been shown to model well a variety of dam break
The numerical model is described in Sect. 2 and flooding problems (STELLING and DUINMEIJER, 2003),
the model-data comparisons are outlined in the sub- and the model formulation appears well suited for the
sections of Sect. 3. The relevance of the benchmarks simulation of long tsunami waves.
to real tsunamis, a preliminary application of the While the non-conservative form of the NLSWEs
model to sediment transport, and the applicability of has no unique solution at local discontinuities, the use
the model are discussed in Sect. 4. Conclusions are of conservative properties, as is done here, is often
given in Sect. 5. sufficient to provide solutions that are acceptable in
terms of the local energy losses in and the propaga-
tion speed of a bore. This is because the conservation
2. Model of mass and momentum should remain valid even for
discontinuities in rapidly varying flows (i.e., breaking
Delft3D is a three-dimensional numerical model tsunami waves) (ZIJLEMA and STELLING, 2008), and
that can simulate coupled hydrodynamic/sediment because the dissipation of energy associated with
transport/morphological change processes. This study wave-breaking-generated turbulence is inherently
examines primarily the hydrodynamic model, which accounted for if momentum is conserved (HIBBERD
solves the nonlinear shallow water equations (NLS- and PEREGRINE, 1979; BROCCHINI and PEREGRINE,
WEs) on a three-dimensional staggered grid using a 1996). However, the assumption of hydrostatic pres-
finite difference scheme (STELLING and VAN KESTER, sure (i.e., a basic assumption of the NLSWEs) is
1994) and the Alternating Direction Implicit time- violated near the break point where vertical acceler-
integration method (STELLING and LEENDERTSE, 1991). ations within the wave are necessary to balance the
The Delft3D model can be run as either a depth- steepening of the wave front. Therefore, models
averaged or vertically layered (i.e., three-dimen- based on the NLSWEs will not capture accurately the
sional) model. In the horizontal, a curvilinear or internal wave dynamics near the point of wave
spherical staggered grid is used. In the vertical (in the breaking. However, once breaking has started, wave
case of three-dimensional simulations) sigma layers, front steepening is balanced by wave-breaking-gen-
which are a fixed percentage of the water depth, or erated turbulence (SVENDSEN and MADSEN, 1984), and
fixed elevation layers are used. When the model is the NLSWEs should be valid for predicting the run-
run in three dimensions, the vertical velocity within up of broken waves (ZIJLEMA and STELLING, 2008).
each grid cell is calculated from the continuity A variety of methods are available to model the
equation due to the hydrostatic assumption and rapid wetting and drying of grid cells (i.e., the
2100 A. Apotsos et al. Pure Appl. Geophys.

moving shoreline) (see ZIJLEMA and STELLING, 2008 efficient manner (i.e., the largest time step that
for a brief discussion of the methods available). Here resulted in no significant change in the results was
a simple method that does not require special drying used). Unless otherwise specified, the cross-shore
and flooding procedures is used (STELLING and DUIN- location (x) is defined as positive onshore from the
MEIJER, 2003), which ensures that the water depth is initial shoreline, the alongshore coordinate (y) is
always positive and requires that only one grid cell positive to the left of the centerline looking onshore,
is flooded or dried per time step. A single time step is the elevation (h) is positive up from the still water
chosen such that this criterion is met throughout each level (SWL), and d is the water depth.
model simulation.
3.1. Run-up on a Planar Beach
3. Numerical Experiments 3.1.1 Setup

The primary goal of this paper is to demonstrate The propagation of an initial free surface disturbance
the general applicability of the model for simulat- composed of a leading depression followed by a
ing tsunami inundation. The model is, therefore, smaller amplitude crest (Fig. 1a, solid grey curve)
compared with a wide array of different experi- over an idealized planar beach with a uniform slope
ments that include several different types of initial of 1:10 (Fig. 1b, solid black line) was solved
wave forms, a wide range of initial nonlinearities, analytically using the initial-value-problem (IVP)
two- and three-dimensional basins, and breaking technique (CARRIER et al., 2003) and the solutions
and non-breaking waves (Table 1). The input were provided by the organizers of the Third
parameters used are not tuned to the observations, International Workshop on Long-Wave Run-up Mod-
but instead physically realistic values were selected els (LIU et al., 2008). Cross-shore snapshots of the
a priori. Owing to the limitations of using labora- free surface elevation (Fig. 2a, solid black curves)
tory experiments and analytical solutions to and depth-averaged velocity (Fig. 2b, solid black
evaluate tsunami inundation models and of using curves) at t = 160, 175, and 220 s, representing the
the NLSWEs to model solitary waves, the model- initial rundown, maximum rundown, and maximum
data comparisons focus on determining if the model run-up, respectively, were provided. Time series of
is generally appropriate for simulating tsunamis the shoreline elevation (Fig. 2c, solid black curve)
(i.e., the model predictions are within *10% of the and velocity (Fig. 2d, solid black curve) from
observations for a variety of numerical and ana- t = 100 to 280 s were also provided.
lytical benchmarks).
For the seven benchmark experiments, the model 3.1.2 Results
is run as a depth-averaged model. Friction is param-
eterized using Manning’s formulation, with Model predictions of the cross-shore water levels and
n = 0.0001 for the analytical solution where friction depth-averaged velocities agree well with the analyt-
is neglected, and n = 0.01, a value consistent with ical solution at all three time steps (Fig. 2a, b
smooth concrete (ARCEMENT and SCHNEIDER, 1990), compare solid black and dashed grey curves). The
for the six laboratory experiments. For each bench- predicted depth-averaged velocity increases sharply
mark, an appropriate time step and grid cell size were near the shoreline for t = 160 and 175 s, possibly
selected to ensure all important wave and bathymetric owing to the depth dependence of the Manning
features were resolved, the stability criterion in the friction formulation. However, this discrepancy is
model that prevents more than one grid cell from confined to water depths less than 0.2 m, and does not
becoming wet/dry during a single time step was sat- affect the model-data agreement in deeper water. The
isfied (i.e., (Dt|u|)/Dx \ 2, where Dt is the time step, predicted shoreline elevation (Fig. 2c) and velocity
u is the flow velocity and Dx is the grid cell size), and (Fig. 2d) are in excellent agreement with the analyt-
the numerical experiments were simulated in an ical solution.
Vol. 168, (2011) Tsunami Model Validation: Toward Sediment Transport 2101

Table 1
Benchmark summary

Benchmark Benchmark Wave form Initial wave Offshore Initial


type height (m) water depth (m) nonlinearity

(1) Planar beach Analytical Leading depression *11 5,000 0.002


(2) Piece-wise linear beach 2-D laboratory Solitary, impulse 0.06–0.54 0.51 0.12–1.06
(3) Alongshore non-uniform beach 3-D laboratory Solitary 0.39 0.78 0.5
(4) Constant depth basin with conical island 3-D laboratory Solitary 0.016, 0.032, 0.064 0.32 0.05, 0.10, 0.20
(5) Alongshore non-uniform beach 3-D laboratory Solitary 0.39 0.78 0.5
with conical island
(6) Complex 3-D bathymetry 3-D laboratory Leading depression *0.025 0.135 0.18
(7) Piece-wise linear beach with vertical wall 2-D laboratory Solitary 0.011, 0.065, 0.153 0.218 0.05, 0.30, 0.70

(a)
5
Water Level (m)

−5

−10

(b)
0
Elevation (m)

−2000

−4000

−50,000 −40,000 −30,000 −20,000 −10,000 0


Distance from Shoreline (m)

Figure 1
a Initial water level disturbance (solid grey curve) and still water level (dashed grey line) and (b) planar bathymetry with a slope of 1:10
(black curve) and initial water level (grey curve)

3.2. Run-up on a Piece-wise Planar Beach gauges and one ADV were located near the limit of
the wave maker during all runs (e.g., Fig. 3, vertical
3.2.1 Setup
hash labeled offshore) to ensure the repeatability and
A laboratory experiment using a two-dimensional alongshore uniformity of the initial wave. The other
piece-wise linear bathymetry was conducted in wave gauges and ADVs were affixed to a movable
the large wave basin at Oregon State University bridge located at x = -3.00, -1.50, -0.75, 0.00,
(BALDOCK et al., 2008) (Fig. 3). Ten different positive 0.25, 0.50, 1.00, 2.00, 3.00 and 6.00 m, respectively
impulse or solitary waves (Table 2, Trial01–Trial10) (Fig. 3, vertical hashes labeled bridge locations). All
were generated in 0.51 m water depth. The simulated instruments were within *0.5 m of the bridge
waves range in nonlinearity (e = Hi/d, where Hi is location. Maximum inundation distances were
the initial wave height) from 0.12 to 1.06, and include recorded visually for all trials. The raw water level
bores, shore breaks, and a non-breaking wave. Water and velocity data were processed to remove spurious
levels were measured using 4 wire resistance wave data (see BALDOCK et al., 2008). However, as this data
gauges and 3 ultrasonic wave gauges. Flow velocities set has not been used extensively, the data were
were measured using 5 Nortek acoustic doppler assessed visually, and time series containing signif-
velocimeters (ADVs). Two wire resistance wave icant noise or non-zero initial water levels were not
2102 A. Apotsos et al. Pure Appl. Geophys.

(a) (c) 20

Water Level (m)


20

Run−up (m)
t = 220 s 10
10
0
0
t = 175 s
−10 −10
t = 160 s
−20 −20
−800 −600 −400 −200 0 200 150 200 250 300

(b)
20 (d) 20
Depth−Averaged

Velocity (m/s)
Velocity (m/s)

t = 175 s 10

Shoreline
10
t = 220 s 0
0
−10
t = 160 s
−10 −20
−800 −600 −400 −200 0 200 150 200 250 300
Distance from Shoreline (m) Time (s)

Figure 2
a Water levels and (b) depth-averaged velocities at t = 160, 175, and 220 s and shoreline (c) elevation (i.e., run-up) and (d) velocity from the
model predictions (dashed grey curves) and the analytical solution (solid black curves). The thick black line in a and b is the bed elevation

0.6
Elevation (m)

0.4 Bridge Locations


Offshore
0.2
d=0.51 m
0
−0.2 1/30
−0.4 1/15

−20 −15 −10 −5 0 5 10 15 20 25


Distance from Shoreline (m)

Figure 3
Model setup for the piece-wise planar beach. Vertical hashes indicate sensor or bridge locations

used. For further description of the model setup and likely because the observed inundation distance,
initial analysis of the results see BALDOCK et al. 15.6 m, occurs at the junction of the onshore slope
(2008). and a horizontal section of the bathymetry (Fig. 3,
x = 15.6 m). A small overestimation of the water
3.2.2 Results velocity in shallow water will cause the wave to
overtop this transition, and flow along the horizontal
The model predicts well (i.e., typically within 10%) bed, resulting in a large change in the predicted
the observed maximum inundation distance for 9 of inundation distance. For both the observations and
the 10 trials (Fig. 4). The model predicts most model predictions the ratio of the maximum run-up
accurately the inundation distance for waves with elevation (R) to the initial wave height (Hi) is largest
small initial nonlinearity, where dispersive effects are for the non-breaking wave, and decreases with
small. While the difference between the model increasing Hi (Table 2), suggesting that wave break-
predictions and observations generally increases with ing reduces the inundation distance of a tsunami.
increasing inundation distance, the model does not The model predicts well the observed wave height
appear to be biased toward over- or underpredicting (within *10%) and wave arrival time (within 1 s) at
the observations. The model significantly overpre- most locations (e.g., Figs. 5 and 6), with the best
dicts the observed inundation distance for Trial01, model-data agreement found for the non-breaking
Vol. 168, (2011) Tsunami Model Validation: Toward Sediment Transport 2103

Table 2
Summary of initial wave conditions for the piece-wise linear beach. Table based on Table 3 in BALDOCK et al. (2008)

Trial # Wave type Hi (m) Surf condition Inundation R (m) Hi/d R/Hi
distance (m)

01 Impulse 0.55 Developed bore 15.6 0.52 1.06 0.95


02 Impulse 0.30 Bore 14.61 0.49 0.58 1.62
03 Impulse 0.17 Shore break 11.35 0.38 0.33 2.23
04 Impulse 0.13 Incipient breaker 10.46 0.35 0.25 2.68
05 Impulse 0.06 Non-breaking 5.97 0.20 0.12 3.32
06 Impulse 0.31 Shore break 11.48 0.38 0.60 1.23
07 Impulse 0.14 Shore break 8.34 0.28 0.27 1.99
08 Solitary 0.1 Incipient breaker 5.78 0.19 0.19 1.93
09 Solitary 0.25 Shore break 9.77 0.33 0.38 1.30
10 Solitary 0.35 Bore 11.83 0.39 0.67 1.13

wave (Fig. 5, Trial05), which has the smallest initial (e.g., Fig. 5, Trial10 and Fig. 6b) or the initial spike
nonlinearity. As the model does not include disper- observed in the water levels during breaking (e.g.,
sion, it does not reproduce accurately the wave shape Fig. 5b, c, Trial10), though the model more accu-
at the offshore locations for highly nonlinear, rately predicts the bore height that follows this spike.
dispersive waves (i.e., large e). For these waves, the The good model-data agreement in terms of water
model predicts that the wave steepens and breaks levels for Trial01, which is highly nonlinear, at all
further offshore than observed. However, the lack of bridge locations (e.g., Fig. 5, Trial01) is likely
dispersion in the model does not significantly affect because all measurements were taken well onshore
the predicted timing of the initial wave, even when of the initial break point. The predicted dissipation in
the wave has broken offshore of the measurement wave energy is owing to the shock-capturing numer-
location (i.e., Fig. 5, Trial01 and Trial10). ical solution scheme used, and not bottom friction.
The model predicts well the wave magnitude in For example, using very low values of n (i.e., a
shallow water and on land, and thus the total decrease virtually frictionless model) does not affect the
in the wave height across the instrumented zone predicted decrease in wave height.
(Fig. 6d, Table 3). However, the model does not The model predictions compare well with the
predict the observed shoaling near the break point measured velocities for most trials and locations (e.g.,
Fig. 7). In shallow water the velocity is well
predicted and increases rapidly as the wave front
passes and then slowly decelerates until the flow is
Predicted Maximum Inundation

16

14 directed offshore (Fig. 7c, d). In deep water the


velocity is less peaked and not as well predicted for
12
Distance (m)

highly nonlinear waves (Fig. 7a, b). Following run-


10
down, differences occur in the observed and
8 predicted time at which sensors become dry. How-
6 ever, this feature is highly sensitive to the value of
4 n used.
2

0
0 2 4 6 8 10 12 14 16 3.3. Run-up on an Alongshore Non-Uniform Beach
Observed Maximum Inundation Distance (m)
3.3.1 Setup
Figure 4
Maximum inundation distance from the observations and model
predictions for Trial02–Trial10. Dashed grey line is perfect
An experimental setup designed to represent a steep,
agreement alongshore, non-uniform continental slope onshore of
2104 A. Apotsos et al. Pure Appl. Geophys.

0.30
0.20 Trial10 Trial01 (a) Bridge at x = −3 m
0.10 Trial05
0
5 10 15 20 25 30 35 40 45 50

0.30
Water Level (m)
0.20 Trial10
Trial01
Trial05
(b) Bridge at x = 0 m
0.10
0
5 10 15 20 25 30 35 40 45 50

0.30
0.20 Trial01 (c) Bridge at x = 1 m
Trial10 Trial05
0.10
0
5 10 15 20 25 30 35 40 45 50

0.30
0.20 (d) Bridge at x = 3 m
Trial10 Trial01
0.10 Trial05
0
5 10 15 20 25 30 35 40 45 50
Time (s)

Figure 5
Water levels for Trial01, Trial05, and Trial10 measured using a resistance wave gauge (solid black curves) and predicted from the model
(dashed grey curves) with the bridge located (a) -3 m, (b) 0 m, (c) 1 m, and (d) 3 m from the initial shoreline

0.6 0.4
(a) x = −17.5 m (b) x = −1.5 m
0.3
0.4
Maximum Predicted Wave Height (m)

0.2

0.2
0.1

0 0
0 0.2 0.4 0.6 0 0.1 0.2 0.3 0.4

0.2
(c) x = 1 m (d) x = 3 m
0.1

0.08

0.1 0.06

0.04

0.02

0 0
0 0.05 0.1 0.15 0.2 0 0.05 0.1
Maximum Observed Wave Height (m)

Figure 6
Maximum predicted and observed wave heights at four cross-shore locations for the resistance wire gauges (squares and diamonds) and sonic
wave gauges (open and filled circles). Dashed grey line is perfect agreement, and positive x is shoreward of the initial shoreline

a deep, constant depth ocean was constructed in the wave 0.39 m high was generated in 0.78 m water
large wave basin at Oregon State University (see depth and allowed to propagate shoreward. Owing to
SWILGER, 2009 for more details) (Fig. 8a). A solitary its large amplitude and high nonlinearity (e = 0.5),
Vol. 168, (2011) Tsunami Model Validation: Toward Sediment Transport 2105

Table 3
Initial wave heights (Hi) and the wave height at x = 3 m for the observations (H3m,obs) and the model predictions (H3m,pred) from the piece-
wise linear beach

Trial 01 02 03 04 05 06 07 08 09 10

Hi (m) 0.55 0.30 0.17 0.13 0.06 0.31 0.14 0.10 0.25 0.35
H3m,obs (m) 0.078 0.08 0.065 0.064 0.047 0.046 0.037 0.029 0.035 0.046
H3m,pred (m) 0.085 0.08 0.072 0.067 0.05 0.046 0.041 0.023 0.039 0.046

2 Trial06 (a) Bridge at x = −3 m


Trial05
0

−2
20 25 30 35 40
Cross−shore Velocity (m/s)

2 Trial06 (b) Bridge at x = 0 m


Trial05
0

−2
20 25 30 35 40

2 Trial06 Trial05 (c) Bridge at x = 1 m


0

−2
20 25 30 35 40

2 Trial06 Trial05 (d) Bridge at x = 3 m


0

−2
20 25 30 35 40
Time (s)

Figure 7
Cross-shore velocities during Trial05 and Trial06 from the observations (solid black curves) and model predictions (dashed grey curves) for
the bridge located (a) -3 m, (b) 0 m, (c) 1 m, and (d) 3 m from the initial shoreline. The oscillations in the observed data occur because the
sensors did not always obtain an accurate measurement, especially during the passage of a breaking wave

the solitary wave broke on the steep shelf slope. 3.3.2 Results
Water levels and velocities were measured at 17 and
3 locations within the basin, respectively (Fig. 8a, The model predicts well (i.e., within 10%) the mag-
white circles). All observations were taken offshore nitude and timing of the initial wave in the constant
of the initial shoreline, and no inundation data were depth portion of the basin (Fig. 9a–b, compare solid
provided. These observations were part of a larger and dashed curves). However, similar to what was
experiment (see SWILGER, 2009), and the data used observed for the piece-wise linear beach, the model
here were provided as part of the ISEC Community does not accurately reproduce the observed wave shape
Workshop: Simulation and Large-Scale Testing of at these locations. As the wave propagates over the
Nearshore Wave Dynamics (http://isec.nacse.org/ steep shelf slope, it becomes more nonlinear and less
workshop/isec_workshop_2009/). dispersive, and the model predicts well the observed
2106 A. Apotsos et al. Pure Appl. Geophys.

Elevation (m)
(a) (b)

Alongshore Distance (m)


10 10 0.2

5 5 0

0 0 −0.2

−0.4
−5 −5

−0.6
−10 −10

−0.8
−20 −10 0 10 −20 −10 0 10
Distance from Shoreline (m)

Figure 8
Model setup for (a) the alongshore non-uniform beach and (b) the alongshore non-uniform beach with a conical island. Open white circles are
the locations where water level measurements are available, and solid white circles are where water level and velocity measurements are
available. In b, only velocity was measured at x = -5 m, y = -5 m. Depth contours are in 0.10 m intervals from -0.80 to 0.40 m. The small
differences observed in the depth contours between (a) and (b) are owing to the fact that the bathymetry provided in (a) is based on a lidar
survey, while the bathymetry in (b) is based on the idealized experimental setup (i.e., no lidar was available)

timing, magnitude and shape of the initial wave 3.4. Run-up on a Circular Island
(Fig. 9c–d). While the model predicts that the wave
3.4.1 Setup
height decreases shoreward owing to breaking, at the
most shoreward locations the maximum predicted A conical island was constructed in the center of a
wave heights are larger (by approximately 40%) and constant depth basin at the US Army Corps of
behind (by approximately 0.5 s) the observations Engineers Waterways Experiment Station, Coastal
(Fig. 9e–f). It is unclear why the model fails to predict Engineering Research Center (LIU et al., 1995;
the total decrease in wave height observed in this BRIGGS et al., 1995; KÁNOGLU and SYNOLAKIS, 1998)
experiment given the good agreement found for the (Fig. 11). The island had a base diameter of 7.2 m
piece-wise linear beach. The discrepancies observed with a slope of 1:4 rising to a height of 0.625 m
here may be owing to the steep slope of the beach and (0.305 m above SWL). As the shoreline is circular,
the proximity to wave breaking. Better model-data the cross-shore coordinate (x) is defined as positive
agreement might be expected further shoreward and on onshore from the wave generating offshore boundary.
land where no observations were taken. Three solitary waves with e = 0.05, 0.1, and 0.2 were
The model predicts well the large cross-shore generated in 0.32 m water depth and propagated
velocities associated with the passage of the initial through the basin. Water levels and run-up are
wave (e.g., Fig. 10a), but does not predict as accu- compared at 4 (Fig. 11, white circles) and 9 locations,
rately the smaller velocities associated with respectively. The values for the run-up were obtained
secondary and reflected waves. Away from the from the benchmark website (http://chl.erdc.usace.
centerline, the model predicts the magnitude of the army.mil/).
observed alongshore velocity, but the predictions lead
the observations for the largest secondary wave 3.4.2 Results
(Fig. 10b, t * 15–20 s). This difference in the
velocities is not related to bed friction (i.e., a larger Consistent with the experiment (LIU et al., 1995), the
n does not improve model-data comparisons), and model predicts the wave refracts around the island.
may indicate the model overpredicts the wave speed The two refracted wave fronts collide in the lee of the
in the very shallow water onshore of the instruments. island, driving flow up the back of the island (i.e., in
Vol. 168, (2011) Tsunami Model Validation: Toward Sediment Transport 2107

0.40
0.20
(a) x = −18.5 m
0

0.40
0.20
(b) x = −14.5 m
0
Water Level (m)

0.40
0.20
(c) x = −13 m
0

0.40
0.20 (d) x = −11 m
0

0.40
0.20
(e) x = −9 m
0

0.40
0.20 (f) x = −5 m
0
0 5 10 15 20 25 30 35 40
Time (s)

Figure 9
Water levels at 6 cross-shore locations along the centerline of the basin for the observations (solid black curves) and model predictions
(dashed grey curves)

(a)
Velocity (cm/s)

150
Cross−shore

100
50
0
−50
0 5 10 15 20 25 30
(b)
Velocity (cm/s)
Alongshore

20

−20

0 5 10 15 20 25 30
Time (s)

Figure 10
(a) Cross-shore and (b) alongshore velocities from the observations (solid black curves) and model predictions (dashed grey curves) for
x = -5 m and y = -5 m

the direction opposite to initial wave propagation). only for e = 0.05 does the predicted wave shape
The model predicts well the timing and magnitude of match the observations. Behind the island, the model
the observed water levels at the two locations in front predicts well the timing of the observed increase in
of the island and the location to the side of the island the water level owing to the arrival of the refracted
(not shown). However, for large e, the front of the waves, but underpredicts the magnitude by about 20,
predicted wave is steeper than the observations and 30 and 50% for e = 0.05, 0.1, and 0.2, respectively.
2108 A. Apotsos et al. Pure Appl. Geophys.

Elevation (m)
(a) 15 0.3

Alongshore Distance (m)


10 0.2

5 0.1

0 0

−0.1
−5

−0.2
−10

−0.3
5 10 15 20 25

(b)
Elevation (m)

0.2
d=0.32 m
0
−0.2
0 5 10 15 20 25
Cross−shore Distance (m)

Figure 11
Model setup for the circular island amid a constant depth basin

The 50% underprediction for e = 0.2 is similar to in 0.78 m water depth and breaks on the steep shelf
that found following TITOV and SYNOLAKIS (1997) (as slope. Water levels and velocities were measured at 9
shown in KÁNOGLU and SYNOLAKIS (1998)) but slightly and 3 locations, respectively (Fig. 8b, white circles).
larger than for the analytical solution presented by Owing to the formation of a strong wake immediately
KÁNOGLU and SYNOLAKIS (1998). behind the island, the sensor located here did not
The model predicts well the variation of run-up accurately record the water level or flow velocity (Pat
around the island for all three waves (Fig. 12). The Lynett, pers. comm.), and these measurements are not
model tends to overpredict the observations for small used. These data were provided as part of the ISEC
e (Fig. 12a) and underpredict the observations for Community Workshop: Simulation and Large-Scale
large e (Fig. 12c). The former may be owing to the Testing of Nearshore Wave Dynamics (http://isec.
predicted oversteepening of the wave in shallow nacse.org/workshop/isec_workshop_2009).
water due to the lack of dispersion in the model,
while the latter is likely caused by the predicted 3.5.2 Results
waves breaking offshore of the island.
The results for the measured water levels and
velocities are similar to those presented in Sect.
3.5. Run-up on an Alongshore Non-Uniform Beach 3.3.2, and are not discussed here. Instead compar-
with a Conical Island isons are made with two features created by the
3.5.1 Setup addition of the island. These include the formation
of a wake in the lee of the island following the
The underlying bathymetry in this experiment is passage of the initial wave (Fig. 13, top panels) and
identical to that described in Sect. 3.3.1 for the the convergence of four secondary waves slightly
alongshore non-uniform beach, except that a conical behind and to the side of the island (Fig. 13, bottom
island was added at the apex of the steep shelf slope panels). Owing to the limited observational data
(Fig. 8b). A solitary wave 0.39 m high was generated available at these locations, the model results are
Vol. 168, (2011) Tsunami Model Validation: Toward Sediment Transport 2109

4 (a) ε = 0.05

0
0 0.5 1 1.5 2 2.5 3

4 (b) ε = 0.1
R/A

0
0 0.5 1 1.5 2 2.5 3

4 (c) ε = 0.2

0
0 0.5 1 1.5 2 2.5 3
Angle from Front of Island (Radians)

Figure 12
Normalized run-up (R/A), where A is the wave amplitude, for the observations (black diamonds) and model predictions (grey circles) for
e = (a) 0.05, (b) 0.1, and (c) 0.2

compared qualitatively with snapshots taken from depression N-wave at the offshore model boundary in
video recordings of the experiments (Pat Lynett, 0.135 m water depth (Fig. 14b). Water level data
pers comm.). were provided at 3 stations inshore of the offshore
The model predicts the formation of a wake island (Fig. 14a). Owing to the complex shoreline,
behind the island, but the shape differs somewhat the cross-shore coordinate (x) for this experiment is
from the observations (Fig. 13, compare top right and positive shoreward from the wave generating off-
left panels). Similarly, the model predicts the con- shore boundary, and the alongshore coordinate (y) is
vergence of four secondary waves, but the magnitude, positive from the right side of the experimental basin
timing, and location of each wave differs somewhat facing onshore.
from the observations (Fig. 13, compare bottom right
and left panels). The slight misalignments of the four 3.6.2 Results
waves likely account for the failure of the model to
capture the signal of this convergence in the water The model predictions are in excellent agreement
level and the velocity time series (not shown) at the with the observations at all three stations for the
single observation location (Fig. 13, black circle in initial wave (Fig. 15). The wave crest arrives at
the bottom left panel). approximately t = 15 s, with the model predictions
lagging the observations by less than 1 s and the
wave magnitude predicted within 10%. After the
3.6. Run-up on a Complex 3-D Beach first wave, the observed water levels are highly
3.6.1 Setup influenced by reflections off the topography and
closed boundaries. The model predicts the arrival of
A 1:400 scale laboratory experiment of the impact of the first reflected wave at approximately t = 35 s,
the 1993 Okushiri tsunami on Monai, Okushiri but the predictions both lead and are larger than the
Island, Japan, including realistic bathymetry and observations. While the model predicts the arrival of
topography scaled to the tank, was conducted in a series of secondary waves at all three locations,
The Central Research Institute for Electric Power the accuracy of the predictions of these reflected
Industry (CRIEPI) in Abiko, Japan (LIU et al., 2008) waves decreases with increasing time (Fig. 15,
(Fig. 14a). The experiment was forced with a leading t [ 55 s).
2110 A. Apotsos et al. Pure Appl. Geophys.

Figure 13
Snapshots of the water surface just after the passage of the initial wave (top left panel) and just as four secondary waves are about to collide
(bottom left panel) and the predicted water levels (color contours) and velocities (vectors) for similar times (right panels). The black circles
and the solid curves in the right panels are the location of an ADV and the depth contours, respectively. Note these panels are rotated 90°
counterclockwise with regards to Fig. 8b

3.7. Run-up on a Vertical Wall near the wall and between gauges 7 and 8, respec-
tively. The wave did not break for e = 0.05.
3.7.1 Setup

A piece-wise linear bathymetry composed of slopes 3.7.2 Results


of 1:53, 1:150, and 1:13 and ending in a vertical wall
was constructed in a 23.2 m long, 0.45 m wide flume For e = 0.05 (case A), the model predicts well the
(BRIGGS et al., 1996) (Fig. 16). Three solitary waves timing, magnitude, and shape of the incoming wave
with e = 0.05, 0.3, and 0.7 were generated in (e.g., Fig. 17a–c). The timing and magnitude of the
0.218 m water depth. Water levels are compared at reflected wave are well predicted by the model, but
6 locations (Fig. 16). The maximum run-up was the predicted wave has a steeper vertical face than
recorded as the highest point water reached on the observed (Fig. 17a, b). This discrepancy in the
vertical wall. For e = 0.3 and 0.7, the waves broke reflected wave shape is because the predicted wave
Vol. 168, (2011) Tsunami Model Validation: Toward Sediment Transport 2111

Elevation (m)
(a) (b) 0.02
0.1
3

Offshore Water Level (m)


Alongshore Distance (m)
0.015

2.5 0.05
0.01
st9
2 0.005
0
st7
1.5 0
st5 −0.05
1 −0.005

0.5 −0.1 −0.01

−0.015
1 2 3 4 5 0 10 20 30
Cross−shore Distance (m) Time (s)

Figure 14
a Model setup and (b) water level at the offshore boundary for the complex 3-D beach. The location of the three water level stations are
labeled st5, st7, and st9 in (a)

0.04
(a) Station 5
0.02
0

10 20 30 40 50 60 70 80 90 100
Water Level (m)

0.04
(b) Station 7
0.02
0
10 20 30 40 50 60 70 80 90 100

(c) Station 9
0.04
0.02
0
10 20 30 40 50 60 70 80 90 100
Time (s)

Figure 15
Water levels at the three stations for the observation (solid black curves) and the model predictions (dashed grey curves)

0.4
Elevation (m)

G10
G7 G8 G9
0.2 G6
G5

d=0.218 m
0

−0.2
−20 −15 −10 −5 0
Distance to Vertical Wall (m)

Figure 16
Model setup for the piece-wise linear beach ending in a vertical wall
2112 A. Apotsos et al. Pure Appl. Geophys.

steepens in the shallow water before reaching the generated spray, which the model cannot predict, as
wall, and maintains this steepened shape after well as other local breaking processes that are not
reflecting seaward. Conversely, the inherent disper- included in the model. Differences in the wave shape
sion of the solitary wave keeps the observed wave as it reaches the wall owing to a lack of dispersion in
from steepening both before and after reflection. The the model may also contribute to the observed
predicted maximum run-up of 0.0246 m is within underprediction.
approximately 10% of the observed maximum of
0.0274 m.
3.8. Kuala Meurisi, Sumatra
For e = 0.3 and 0.7 (cases B and C, respectively),
dispersion becomes more important, and the model 3.8.1 Setup
predicts well the timing, amplitude, and shape of the
incoming wave only at station 5 (Fig. 17d, g). Further Bathymetry, topography, and maximum water level
shoreward the model predicts that the wave steepens estimates were collected following the 26 December
and breaks at a location more offshore than observed. 2004 Indian Ocean tsunami near Kuala Meurisi on
Significant differences exist between the modeled and the north coast of Sumatra, Indonesia (JAFFE et al.,
observed reflected waves in terms of both the wave 2006; GELFENBAUM et al., 2007) (Fig. 18a). The
magnitude and shape, likely owing to the way a maximum water levels were mostly estimated from
broken or breaking wave interacts with a vertical broken branches, and may represent minimum values
surface. While the timing of the reflected wave is of the actual water levels. The bathymetry was
generally well predicted for e = 0.3, for e = 0.7 the measured along seven cross-shore transects, while
predictions lead the observations by about 1 s at the topography was measured only along a single tran-
most seaward location (Fig. 17g). The predicted sect. As the bathymetry is approximately alongshore
maximum run-up values of 0.0936 m and 0.154 m uniform, a one-dimensional, depth-averaged model
for e = 0.3 and 0.7, respectively, are about 4 and 2 with a cross-shore grid spacing of 12 m is used. The
times smaller than the observed values of 0.4572 m offshore water level variation was obtained from the
and 0.2743 m. This underprediction is likely owing to output of a deep water ocean propagation model and
the measured run-up being a result of wave-breaking- an estimate of the initial source disturbance for the 26

Case A Case B Case C


(a) Station 5 (g) Station 5
0.015
0.08 (d) Station 5
0.06 0.1
0.01 0.04
0.005 0.02 0.05
0 0 0
5 10 15 20 5 10 15 20 5 10 15
Water Level (m)

(b) Station 8 0.08 (e) Station 8 (h) Station 8


0.015 0.06 0.1
0.01 0.04
0.005 0.02 0.05
0 0 0
5 10 15 20 5 10 15 20 5 10 15

(c) Station 10 (f) Station 10 (i) Station 10


0.08
0.015 0.06 0.1
0.01 0.04
0.005 0.02 0.05
0 0 0
5 10 15 20 5 10 15 20 5 10 15
Time (s)

Figure 17
Water levels for e = (a–c) 0.05, (d–f) 0.3, and (g–i) 0.7 at station (a, d, g) 5, (b, e, h) 8, and (c, f, i) 10 for the model (solid black curves) and
observations (dashed grey curves). In each panel, the first and second waves are the incident and reflected waves, respectively
Vol. 168, (2011) Tsunami Model Validation: Toward Sediment Transport 2113

(a) 20

Elevation (m)
10

−10

−2000 −1500 −1000 −500 0 500 1000 1500 2000


Distance from Shoreline (m)
(b)
Peaked Peaked
Level (m)

5 Wave
Wave
Water

Broad
0
Wave
−5
0 20 40 60 80 100 120
Time (min)

Figure 18
a Observed (black circles) and predicted (dashed black curve) maximum water levels at Kuala Meurisi. The solid black curve is the combined
bathymetry and topography and the shaded grey area represents the range of maximum water levels predicted for the three simulations
conducted. b Offshore wave level shoaled to 17 m water depth

December 2004 tsunami (VATVANI et al., 2005a, b) onshore. A larger onshore roughness is used to take
(Fig. 18b). Due to the coarse resolution of the deep into account the presence of vegetation.
water model, the offshore water level variation was
available only in 35 m water depth. As the high 3.8.2 Results
resolution nearshore bathymetry used in this study
extends only to 17 m water depth (Fig. 18a), The model predicts the run-up elevation well, but
Green’s law (H * h1/4) is used to shoal the overpredicts the observed water levels closest to the
predicted waves from 35 to 17 m water depth. This shoreline while underpredicting the water levels
shoaling produces approximately a 20% increase in further shoreward (Fig. 18a, compare dashed black
the wave height of the water level variation used to curve and open circles). The observed overprediction
force the model simulation. Changes in the wave near the shoreline may be owing to biased estimates
length are inherently included by forcing the model of the observed water levels, as no large trees
with a time series, given that the wave period remained standing after the tsunami. Given the large
remains relatively constant. However, no attempt uncertainty in the model setup, including in the
was made to simulate the change in wave shape due offshore water level and bed roughness, the assump-
to shoaling. The entire wave train, which is tion of alongshore uniformity, the use of bathymetry
composed of three large waves (i.e., two peaked taken after the tsunami, and the exclusion of local
waves and a broad wave) followed by a series of obstacles such as houses and trees, the model predicts
smaller waves, is simulated. This wave forcing is a the observations surprisingly well.
simplification of the actual tsunami as it is assumed As a first step to quantify the effect of the
to be normally incident, and therefore neglects uncertainty on the maximum water level predictions,
waves propagating at angles to the coast, such as two further simulations were conducted. The first was
trapped and scattered waves, and may not capture forced using the unshoaled water level variation
accurately all wave reflections off nearby land predicted in 35 m water depth (i.e., an assumed lower
masses. Bed roughness is estimated using Manning’s bound on the wave forcing). The second was forced
formulation with n = 0.025 and 0.032 off- and using a 40% increase in the water level from 35 m
2114 A. Apotsos et al. Pure Appl. Geophys.

water depth coupled with a 50% decrease in the on- hydrodynamics, usually only highly nonlinear, and
and offshore roughness (i.e., an assumed upper thus relatively short, solitary waves are simulated.
bound). These simulations indicate that while the For example, in the laboratory studies examined here,
uncertainty has an effect on the predictions, the the nonlinearity of the solitary waves varies from
predictions from all three simulations fall within the 0.05 to 0.67. However, the nonlinearity of actual
range of the observations (Fig. 18a, compare shaded tsunamis observed away from the nearshore is
grey area and black circles). Future work will typically much smaller than this, and previous studies
examine in more detail the sensitivity of the flow (MADSEN et al., 2008) have suggested that solitary
predictions to the input conditions. waves with e [ 0.05 are not long waves, and are
more similar to wind waves than to tsunamis.
Furthermore, observations and eyewitness
4. Discussion accounts from several recent tsunamis indicate that
the largest wave does not always arrive first (e.g.,
4.1. Solitary Waves as Tsunamis PAPADOPOULOS et al., 2006; CHOOWONG et al., 2008)
and that the first wave can arrive as a leading
Over the past few decades many analytical and
depression (TADEPALLI and SYNOLAKIS, 1996). Both of
laboratory studies of tsunami propagation and inun-
these observations argue against the development of
dation have used solitary waves as proxies for
solitons, which requires the largest wave to arrive
tsunami waves. It was first suggested by HAMMACK
first and be of positive orientation. Some studies have
(1973), SEGUR (1973), and HAMMACK and SEGUR
suggested that leading depression N-waves are a
(1974, 1978a, b) that a positive initial surface
better representation of tsunamis (TADEPALLI and
disturbance of arbitrary shape (i.e., an initial tsunami
SYNOLAKIS, 1994, 1996). However, these waves are
wave generated in deep water) will eventually
difficult to generate in the laboratory and have not
devolve into a series of solitary waves if the
been widely used in experimental studies. Further-
propagation distance is long enough. Based on this
more, the use of single waves, whether of solitary or
observation and the fact that solitary waves are easily
N shape, will not produce wave–wave interactions,
reproducible in the laboratory, are defined by only
which may be important for accurately predicting
two parameters (i.e., Hi and d), and can be modeled in
both the maximum inundation distance and sediment
the limited spatial scales of most laboratory facilities,
transport of tsunamis.
solitary waves have continued to be used extensively
Unfortunately, the limited horizontal dimensions
to model tsunamis, including in 5 of the 7 bench-
of laboratory facilities and the difficulty of generating
marks examined in this study.
more complex wave forms will limit the nonlinearity,
Recently, however, a growing number of authors
number, and shape of the waves that can be
have questioned the relevance of solitary waves as
generated. Therefore, until more realistic experiments
proxies for tsunamis (MADSEN et al., 2008; CONSTAN-
can be conducted or more detailed observations from
TIN and JOHNSON, 2008; CONSTANTIN, 2009). Using
actual tsunamis recorded, these laboratory experi-
geophysically realistic scales of nonlinearity and
ments will remain an important aspect of validating
dispersion, these studies demonstrate that the distance
tsunami propagation and inundation models.
necessary for earthquake-generated tsunami waves to
devolve into solitons can be several times the fetch of
the largest ocean basin. Further limiting the relevance
4.2. Scaling Up to Sediment Transport
of solitary waves as proxies for tsunamis is the fact
that their nonlinearity and length are not independent. Extending numerical models to include sediment
Instead, the relative length of a solitary wave transport increases the complexity and uncertainty of
decreases with increasing nonlinearity. Owing to the the model. For example, near bed fluid velocities and
limited spatial dimensions of most laboratory facil- the suspension of sediment are highly sensitive to the
ities and the need to record measurable bed roughness, which may vary significantly
Vol. 168, (2011) Tsunami Model Validation: Toward Sediment Transport 2115

spatially. Calculating sediment transport also requires domain. While the erodable bed thickness at this site
the delineation of the sediment source characteristics is unknown, the maximum predicted scour is less
(i.e., location, grain-size distribution, thickness), than 5 m, and therefore the onshore deposition of
which are often poorly constrained, as well as the sediment is not limited by the thickness used.
use of transport formulations developed empirically Sediment transport is calculated by solving the
for lower velocity flows. conservation of mass equations within the hydrody-
Unfortunately, no standardized benchmarks cur- namic model. Bedload and suspended load transport
rently exist with which to verify model predictions of are calculated following VAN RIJN (1993; 2007a, b)
tsunami-induced sediment transport, though preli- combined with a k-e turbulence model. Bedload is
minary work conducted as part of a workshop held in calculated based on a sediment mobility number and
Friday Harbor, Washington, in 2007 has shown some the critical velocity for the initiation of motion.
initial success (HUNTINGTON et al., 2007). The devel- Suspended load is calculated based on a near-bed
opment of laboratory-scale benchmarks is not reference concentration and a vertical diffusion
straightforward because these benchmarks must coefficient determined from the output of the turbu-
include and balance the effects of sediment suspen- lence model. Sediment fluxes between the water
sion, advection, and settling on the appropriate time column and the bed are calculated within each grid
and length scales. This is difficult given the limited cell based on the vertical diffusion of sediment owing
size of most laboratory facilities and the fact that to turbulent mixing and deposition from settling, and
sediment becomes cohesive at small grain sizes. The the bottom morphology is updated accordingly at
use of solitary waves, which dominate laboratory each time step (LESSER et al., 2004; VAN RIJN et al.,
studies, may not be appropriate because the flow 2004). Sediment-induced density stratification owing
velocities and accelerations induced are likely to be to high suspended sediment concentrations is inher-
different than those induced by long tsunami waves. ently accounted for within the turbulence model by
Development of sediment transport benchmarks adjusting the fluid density to include the mass of the
based on field observations of sediment deposits suspended sediment. The settling velocity is deter-
collected after a tsunami is hindered by the fact that mined based on the clear water settling velocity of
the initial conditions (i.e., the nearshore wave form, VAN RIJN (1993) combined with the effects of
sediment source characteristics) and onshore rough- hindered settling following RICHARDSON and ZAKI
ness, especially in the presence of vegetation, are (1954). The combined sediment transport/morpho-
typically poorly constrained. Furthermore, even good logical change model has been shown to model well
agreement with observed water levels (i.e., Sect. many coastal and estuarine settings (LESSER et al.,
3.8.2), often the only hydrodynamic information 2004; GERRITSEN et al., 2007; VAN RIJN et al., 2007).
available, does not necessary imply good agreement Using this simplified setup, the model predicts the
with the time-varying velocities, which dictate the observed sediment deposit thickness generally within
suspension and transport of sediment, that occurred a factor of 2, though locally the predictions can be off
during the inundating tsunami. by more than an order of magnitude (Fig. 19,
The difficulties associated with modeling tsu- compare dashed grey curve with black circles). Both
nami-induced sediment transport not-withstanding, model simulations and field measurements show
the Kuala Meurisi simulation conducted in Sect. 3.8 thicker deposits in local topographic lows and thinner
was extended to include the suspension and transport deposits or erosion on local topographic highs. Most
of sediment. First, the water column was subdivided of the sediment is deposited by the second peaked
into ten vertical sigma layers (i.e., layers representing wave, although a smaller fraction is deposited by the
a constant percentage of the water column), with first peaked wave. Very little sediment was eroded or
increasing resolution near the bed, to accurately deposited by the broad wave. The observed peak in
represent the bottom boundary layer. An erodable sediment thickness near the cliff is composed of fine
sediment bed 5 m thick, with a mean grain size of sediment and is not captured by the model which uses
400 lm was assumed to exist over the entire model only a single grain size. Given the high level of
2116 A. Apotsos et al. Pure Appl. Geophys.

Thickness (m)
0.4

Deposit
0.2

0 200 400 600 800 1000 1200 1400 1600 1800 2000
Distance from Shoreline (m)

Figure 19
Deposit thickness measured near Kuala Meurisi (black circles) and change in bed elevation predicted by the model (dashed grey curve)

uncertainty involved in the model setup, the model- As stated previously, comparisons with laboratory
data agreement is remarkably good. More work needs observations offer only qualified validation of model
to be done to examine the details of the predicted applicability. Therefore, while the model predictions
sediment transport and deposition to ensure that the are in reasonable agreement with benchmarks in
model is accurately capturing the appropriate pro- which dispersion is important (i.e., highly nonlinear
cesses. However, the detailed analyses required are solitary waves), this does not necessarily imply the
beyond the scope of this paper, and will be the subject model is appropriate for modeling highly dispersive
of future work. tsunamis. Tsunamis generated by landslides, volcanic
eruptions, and meteoroid impacts can have wave-
lengths of the same order of magnitude as the water
4.3. Model Applicability
depth, and frequency dispersion can be important
Delft3D predicts well the general hydrodynamic even near the source. A previous study (LYNETT et al.,
observations associated with tsunami nearshore prop- 2003) found large differences between offshore wave
agation and inundation measured under a wide range height predictions from models based on the NLS-
of conditions. This study, combined with previous WEs and on the Boussinesq equations, which allow
studies (VATVANI et al., 2005a, b; GELFENBAUM et al., for weak frequency dispersion, for the landslide-
2007) that showed Delft3D models well the deep- generated component of the 1998 Papua New Guinea
water propagation and inundation of the 26 December tsunami. Therefore it is unclear how accurately
2004 Indian Ocean tsunami near Banda Aceh, Delft3D will model the deep-water propagation of
demonstrates that Delft3D is applicable for modeling highly dispersive waves such as those generated by a
long, non-dispersive tsunamis from the source to the landslide.
maximum extent of inundation. The good model-data agreement demonstrates the
As Delft3D is based on the NLSWEs, which are numerical solution method employed is appropriate
derived using the assumption of hydrostatic pressure, for modeling the general characteristics (i.e., wave
the model does not include dispersive terms and is height, run-up, velocities) of both breaking and non-
best suited to tsunamis where the wavelength is long breaking waves. The advantage of this numerical
compared to the water depth. The model-data com- solution technique over many Boussinesq models is
parisons presented here support this, with the model that it does not require a breaking criterion, but
predicting most accurately the observations from inherently captures the loss of energy associated with
benchmarks where the wave is long (i.e., the analyt- wave breaking. Furthermore, the accuracy of the
ical solution over a planar beach) and where predictions onshore of breaking does not appear to
dispersion is relatively less important (i.e., Trial05 decrease for waves with higher initial nonlinearity
from the piece-wise linear experiment). Therefore, (e.g., Fig. 5d), which may suggest depth-limited
Delft3D is likely appropriate for modeling most breaking masks the effects of frequency dispersion
earthquake-induced tsunamis, which normally have onshore of breaking. A previous study (LYNETT et al.,
very long wavelengths and experience negligible 2003) found similar nearshore wave height predic-
frequency dispersion over typical propagation tions using both an NLSWE and a Boussinesq model
distances. for the landslide-generated component of the 1998
Vol. 168, (2011) Tsunami Model Validation: Toward Sediment Transport 2117

Papua New Guinea tsunami even though the model 5. Conclusions


predictions further offshore were significantly differ-
ent. It was suggested that the similarity in the The model predicts well the hydrodynamic
nearshore wave height predictions was owing to observations (i.e., water levels, velocities, and inun-
depth-limited breaking. While this may imply that dation distances) from seven analytical and
Delft3D can be used to model the run-up and laboratory experiments and the maximum water lev-
inundation of breaking, dispersive waves, under els observed near Kuala Meurisi, Sumatra, following
exactly what conditions the NLSWEs accurately the 26 December 2004 tsunami. Model predictions
represent the hydrodynamics of dispersive tsunamis agree best with experiments where dispersion is less
onshore of breaking remains to be determined. important. The model predicts well the total decrease
While the model predicts well the overall dissi- in wave height and the speed of breaking waves.
pation of wave energy owing to breaking, the model However, the model does not reproduce the observed
does not predict local breaking dynamics such as the shoaling for some waves or the shape of highly
generation of turbulence or vertical flow accelera- nonlinear waves in deep water. These results indicate
tions. Even for waves that are non-dispersive in deep the model is appropriate for simulating tsunami
water, dispersion becomes important close to break- propagation and inundation for both breaking and
ing where it balances the increasing nonlinearity non-breaking long waves (i.e., most earthquake-
associated with shoaling and delays breaking (CONST- induced tsunamis) but may not be appropriate for
ANTIN and JOHNSON, 2008). As the model does not modeling highly dispersive tsunamis, such as those
include dispersion, it will predict that waves break generated by landslides or impacts, especially over
further offshore than physically realistic. Therefore, long distances.
while the model may predict well the water levels and Extension of the model to include sediment
velocities before and after breaking, as well as the transport is shown to produce reasonable results for
maximum run-up elevations of breaking waves, it is long, non-breaking waves, but may not be appropriate
not appropriate for simulating wave characteristics in the case of breaking or dispersive waves. While
near the breakpoint. Furthermore, the model may including sediment transport increases the complexity
underpredict the run-up and inundation of waves of and uncertainty in the model, the model predicts
close to breaking, as the lack of dispersion will cause the sediment deposition at Kuala Meurisi generally
these otherwise non-breaking waves to break and within a factor of two. These results suggest that the
dissipate energy. coupling of validated tsunami-induced sediment
The generally good agreement with the observed transport models with paleo-tsunami deposits could
water levels and flow velocities suggests extension aid in the improvement of local tsunami hazards
of the model to include sediment transport may be assessments.
appropriate for long, non-breaking waves generated
by earthquakes. However, extension of the model
may not be appropriate for dispersive or breaking Acknowledgements
waves. In the case of dispersive waves, the wave
shape, and thus the cross-shore distribution of This research was funded by a USGS Mendenhall
velocity, is poorly predicted, and the model may Post-Doctoral Fellowship and the USGS Coastal
misrepresent the distribution of the bed stress. For and Marine Geology Program. Reviews provided
breaking waves, the location and magnitude of by Eric Geist, Pat Lynett, Utku Kânoğlu, and 2
wave-breaking-generated turbulence, which the anonymous reviewers greatly improved this manu-
model does not predict for long waves, will affect script. Edwin Elias is thanked for his help and
the patterns of sediment suspension, transport, and guidance concerning the details of the numerical
deposition. model.
2118 A. Apotsos et al. Pure Appl. Geophys.

REFERENCES HAMMACK, J. (1973), A note on tsunamis: Their generation and


propagation in an ocean of uniform depth, J. Fluid Mech., 60(4),
769-799.
ARCAS, D. and AITOV, V. (2006), Sumatra tsunami: lessons from
HAMMACK, J. and SEGUR, H. (1974), The Korteweg-deVries equation
modeling, Surv. Geophys., 27, 679-705.
and water waves, part 2: Comparison with experiments, J. Fluid
ARCEMENT, G. and SCHNEIDER, V. (1990), Guide for selecting
Mech., 65(2), 289-314.
Manning’s roughness coefficients for natural channels and flood
HAMMACK, J. and SEGUR, H. (1978a), The Korteweg-deVries equa-
plains, Water Supply Pap., U.S. Geol. Surv., Reston, VA.
tion and water waves, part 3: Oscillatory waves, J. Fluid Mech.,
ATWATER, B.F. (1987), Evidence for great Holocene earthquakes
84(2), 337-358.
along the outer coast of Washington State, Science, 236,
HAMMACK, J. and SEGUR, H. (1978b), Modeling criteria for long
942-944.
waves, J. Fluid Mech., 84(2), 359-373.
BALDOCK, T.E., COX, D., MADDUX, T., KILLIAN, J., and FAYLER, L.
HIBBERD, S. and PEREGRINE, D. (1979), Surf and run-up on a beach:
(2008), Kinematics of breaking tsunami wave fronts: A dataset
a uniform bore, J. Fluid Mech., 95, 323-345.
from large scale laboratory experiments, Coastal Eng., doi:
HUNTINGTON, K., BOURGEOIS, J., GELFENBAUM, G., LYNETT, P., JAFFE,
10.1016/j.coastaleng.2008.10.011.
B., YEH, H., and WEISS, R. (2007), Sandy signs of a tsunami’s
BRIGGS, M.J., SYNOLAKIS, C.E., HARKINS, G.S., and GREEN, D.R.
onshore depth and speed, EOS Trans., 88(52), 577-584.
(1995), Laboratory experiments of tsunami run-up on a circular
JAFFE, B. and GELFENBAUM, G. (2007), A simple model for calcu-
island, Pure Appl. Geophys., 144, 569-593.
lating tsunami flow speed from tsunami deposits, Sediment.
BRIGGS, M.J., SYNOLAKIS, C.E., KANOGLU, U., and GREEN, D.R.
Geol., 200, 347-361.
(1996), Benchmark Problem #3: Runup of solitary waves on a
JAFFE, B., BORRERO, J., PRASETYA, G., PETERS, R., MCADOO, B.,
vertical wall, ‘‘Long-wave Runup Models,’’ International
GELFENBAUM, G., MORTON, R., RUGGIERO, P., HIGMAN, B., DEN-
Workshop on Long Wave Modeling of Tsunami Runup, Friday
GLER, L., HIDAYAT, R., KINGSLEY, E., KONGKO, W., LUKIJANTO,
Harbor, San Juan Island, WA, September 12-16, 1995.
MOORE, A., TITOV, V., and YULIANTO, E. (2006), Northwest
BROCCHINI, M. and PEREGRINE, D. (1996), Integral flow properties of
Sumatra and offshore islands field survey after the December
the swash zone and averaging, J. Fluid Mech., 317, 241-273.
2004 Indian Ocean tsunami, Earthquake Spectra, 22, 105-135.
CARRIER, G.F. and GREENSPAN, H.P. (1958), Water waves of finite
KÁNOĞLU, U. (2004), Nonlinear evolution and runup-rundown of
amplitude on a sloping beach, J. Fluid Mech., 4, 97-109.
long waves over a sloping beach, J. Fluid Mech., 513, 363-372.
CARRIER, G.F., WU, T.T., and YEH, H. (2003), Tsunami run-up and
KÁNOĞLU, U. and SYNOLAKIS, C.E. (1998), Long wave runup on
draw-down on a plane beach, J. Fluid Mech., 475, 79-99.
piecewise linear topographies, J. Fluid Mech., 374, 1-28.
CHOOWONG, M., MURAKOSHI, N., HISADA, K., CHARUSIRI, P.,
LIU, P.L.-F., CHO, Y.S., BRIGGS, M.J., KANOGLU, U., and SYNOLAKIS,
CHAROENTITIRAT, T., CHUTAKOSITKANON, V., JANKAEW, K.,
C.E. (1995), Run-up of solitary waves on a circular island, J.
KANJANAPAYONT, P., and PHANTUWONGRAJ, S. (2008), 2004 Indian
Fluid. Mech., 302, 259-285.
Ocean tsunami inflow and outflow at Phuket, Thailand, Mar.
LIU, P.L.-F., YEH, H., and SYNOLAKIS, C.E. (ed.) (2008), Advanced
Geo., 248, 179-192.
numerical models for simulating tsunami waves and run-up,
CONSTANTIN, A. (2009), On the relevance of soliton theory to tsu-
Advances in Coastal and Ocean Engineering, 10.
nami modeling, Wave Motion, 6, 420-426.
LESSER, G.R., ROELVINK, J.A., VAN KESTER, J.A.T.M., and STELLING,
CONSTANTIN, A. and JOHNSON, R.S. (2008), Propagation of very long
G.S. (2004), Development and validation of a three-dimensional
waves, with vorticity, over variable depth, with applications to
morphological model, Coastal Eng., 51, 883-915.
tsunamis, Fluid Dyn. Res., 40, 175-211.
LYNETT, P., BORRERO, J., LIU, P., and SYNOLAKIS, C.E. (2003), Field
FRITZ, H., BORRERO, J., SYNOLAKIS, C., and YOO, J. (2006), 2004
survey and numerical simulations: A review of the 1998 Papua
Indian Ocean flow velocity measurements from survivor videos,
New Guinea Tsunami, Pure Appl. Geophys., 160, 2119-2146.
Geophys. Res. Lett., 33, L24605.
MADSEN, P., FUHRMAN, D., and SCHAFFER, H. (2008), On the solitary
GELFENBAUM, G., VATVANI, D., JAFFE, B., and DEKKER, F. (2007),
wave paradigm for tsunamis, J. Geophys. Res., 113, C12012, doi:
Tsunami inundation and sediment transport in the vicinity of
10.1029/2008JC004932.
coastal mangrove forest, Conference Proceedings Coastal Sedi-
NANAYAMA, F., FURUKAWA, R., SHIGENO, K., MAKINO, A., SOEDA, Y.,
ment ‘07, doi: 10.1061/40926(239)86.
and IGARASHI, Y. (2007), Nine unusually large tsunami deposits
GERRITSEN, H., DE GOEDE, E.D., PLATZEK, F.W., GENSEBERGER, M.,
from the past 4000 years at Kiritappu marsh along the southern
VAN KESTER, J.A.TH.M., and UITTENBOGAARD, R.E. (2007), Vali-
Kuril Trench, Sediment. Geol., 200, 275-294.
dation document for Delft3D-Flow, Rep. X0356, Delft
PAPADOPOULOS, G., CAPUTO, R., MCADOO, B., PAVLIDES, S., KA-
Hydraulics, Delft, The Netherlands.
RATATHIS, V., FOKAEFS, A., ORFANOGIANNAKI, K., and VALKANIOTIS,
GONZALEZ, F., BERNARD, E., MEINIG, C., EBLE, M., MOFJELD, H., and
S. (2006), The large tsunami of 26 December 2004: Field
STALIN, S. (2005), The NTHMP Tsunameter network, Nat.
observations and eyewitnesses accounts from Sri Lanka, Mal-
Hazards, 35, 25-29.
dives Is. and Thailand, Earth, Planets, Space, 58, 233-241.
GONZÁLEZ, F.I., GEIST, E.L., JAFFE, B., KÂNOĞLU, U., MOFJELD, H.,
RICHARDSON, J.F. and ZAKI, W.N. (1954), Sedimentation and fluid-
SYNOLAKIS, C.E., TITOV, V.V., ARCAS, D., BELLOMO, D., CARLTON,
ization: Part I, Trans. Inst. Chem. Eng., 32, 35-50.
D., HORNING, T., JOHNSON, J., NEWMAN, J., PARSONS, T., PETERS, R.,
SEGUR, H. (1973), The Korteweg-deVries equation and water
PETERSON, C., PRIEST, G., VENTURATO, A., WEBER, J., WONG, F.,
waves, part 1: Solutions of the equations, J. Fluid Mech., 59(4),
and YALCINER, A., (2009), Tsunami pilot study working group
721-735.
‘‘Probabilistic tsunami hazard assessment at Seaside, Oregon for
SMITH, D.E., FOSTER, I.D.L., LONG, D., and SHI, S. (2007), Recon-
near- and far-field seismic sources,’’ J. Geophys. Res., 114,
structing the pattern and depth of flow onshore in a paleo-
C11023, doi:10.1029/2008JC005132.
Vol. 168, (2011) Tsunami Model Validation: Toward Sediment Transport 2119

tsunami from associated deposits, Sediment. Geol., 200, TITOV, V., GONZALEZ, F., BERNARD, E., EBLE, M., MOFJELD, H.,
362-371. NEWMAN, J., and VENTURATO, A., (2005), Real-time tsunami
STELLING, G.S. and DUINMEIJER, S.P.A. (2003), A numerical method forecasting, Natural Hazards, 35(1), 45-58.
for every Froude number in shallow water flows, including large U.S. Agency for International Development (USAID) (2005), Fact
scale inundations, Int. J. Num. Meth. in Fluids, 43, 1329-1354. sheet, July 7. http://www.usaid.gov/our_work/humanitarian_
STELLING, G.S. and LEENDERTSE, J.J. (1991), Approximation of assistance/disaster_assistance/countries/Indian_ocean/fy2005/
convective processes by cyclic ACI methods. In: Proceedings of indianocean_et_fs39-07-07-2005.pdf.
2nd ASCE Conference on Estuarine and Coastal Modeling, VAN RIJN, L.C. (1993), Principles of sediment transport in rivers,
Tampa, USA, 771–782. estuaries and coastal seas, Aqua Publications, Amsterdam.
STELLING, G.S. and VAN KESTER, J.A.T.M. (1994), On the approxi- VAN RIJN, L.C. (2007a), Unified view of sediment transport by
mation of horizontal gradients in sigma coordinates for currents and waves. I: Initiation of motion, bed roughness, and
bathymetry with steep bottom slopes, Int. J. Num. Meth. in bed-load transport, J. Hydraul. Eng., 133, 649-667.
Fluids, 18, 915-935. VAN RIJN, L.C. (2007b), Unified view of sediment transport by
SVENDSEN, I.A. and MADSEN, P.A. (1984), A turbulent bore on a currents and waves. II: Suspended transport, J. Hydraul. Eng.,
beach, J. Fluid Mech., 148, 73-96. 133, 668-689.
SWILGER, D.T. (2009), Laboratory study investigating the three VAN RIJN, L.C., WALSTRA, D.J.R., and VAN ORMONDT, M. (2004),
dimensional turbulence and kinematic properties associated with ‘‘Description of TRANSPOR 2004 (TR2004) and implementation
a breaking solitary wave, Master Thesis, Texas A&M. in DELFT3D-online,’’ Rep. Z3748, Delft Hydraulics, Delft, The
SYNOLAKIS, C.E. (1987), The run-up of solitary waves, J. Fluid Netherlands.
Mech., 185, 523-545. VAN RIJN, L.C., WALSTRA, D., and VAN ORMONDT, M. (2007), Unified
SYNOLAKIS, C.E., BERNARD, E.N., TITOV, V.V., KANOGLU, U., and view of sediment transport by currents and waves. IV: Applica-
GONZALEZ, F. (2007), Standards, criteria, and procedures for tion of morphodynamic model, J. Hydraul. Eng., 133, 776-793.
NOAA evaluation of tsunami numerical models, NOAA OAR VATVANI, D., SCHRAMA, E., and VAN KESTER, J. (2005a), Hindcast of
Special Report, Contribution No 3053, NOAA/OAR/PMEL, tsunami flooding in Aceh-Sumatra, Proc. of the 5th International
Seattle, Washington. Symposium on Ocean Wave Measurement and Analysis, Madrid,
SYNOLAKIS, C.E., BERNARD, E.N., TITOV, V.V., KANOGLU, U., and Spain.
GONZALEZ, F. (2008), Validation and verification of tsunami VATVANI, D., BOON, J., and RAMANAMURTY, P.V. (2005b), Flood risk
numerical models, Pure Appl. Geophys., 165, 2197-2228. due to tsunami and tropical cyclones and the effect of tsunami
TADEPALLI, S. and SYNOLAKIS, C.E. (1994), The run-up of N-waves excitations on tsunami propagations, Proc. of the IAEA Work-
on sloping beaches, Proc. R. Soc. Lond., 445, 99-112. shop on External Flooding Hazards, Tamil Nadu, India.
TADEPALLI, S. and SYNOLAKIS, C.E. (1996), Model for the leading WEI, Y., BERNARD, E., TANG, L., WEISS, R., TITOV, V., MOORE, C.,
waves of tsunamis, Phys. Rev. Lett., 2141-2144. SPILLANE, M., HOPKINS, M., and KÁNOĞLU, U. (2008), Real-time
TANG, L., TITOV, V., and CHAMBERLIN, C. (2009), Development, experimental forecast of the Peruvian tsunami of August 2007 for
testing, and application of site-specific tsunami inundation U.S. coastlines, Geophys. Res. Letters, 35, L04609.
models for real-time forecasting, J. Geophys. Res., 114, C12025, YEH, H., LIU, P., and SYNOLAKIS, C.E., (ed.) (1996), Proc. Int.
doi:10.1029/2009JC005476. Symposium, Friday Harbor, USA, 12–17 September 1995, Long-
TITOV, V. and SYNOLAKIS, C.E. (1997) Numerical modeling of 2-D Wave Rump Models, World Sci., Singapore.
and 3-D long wave runup using VTSC-2 and VTSC-3. In Long ZIJLEMA, M. and STELLING, G.S. (2008), Efficient computation of
wave runup models (ed. H. Yeh et al.), World Scientific, surf zone waves using the nonlinear shallow water equations
242-248. with non-hydrostatic pressure, Coastal Eng., 55, 780-790.

(Received January 21, 2010, revised December 8, 2010, accepted December 19, 2010, Published online March 4, 2011)

You might also like